Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Dressel.Gruner.Electrodynamics of Solids.2003

.pdf
Скачиваний:
66
Добавлен:
27.08.2013
Размер:
3.91 Mб
Скачать

108

5 Metals

Here E and H are uniform, frequency dependent electric and magnetic fields. Finally, collisions may cause a transition from one state |k to another |k :

 

f

scatter = [ fk (1fk)fk(1fk )]W (k, k ) dk = [ fk fk]W (k, k ) dk .

 

t

 

 

 

 

 

 

 

(5.2.6)

In this

equation W (k, k ) represents the transition probability between the initial

state |k and the final state |k with the corresponding wavevectors; we assume microscopic reversibility W (k, k ) = W (k , k). Here fk denotes the number of carriers in the state |k , and 1 fk refers to the number of unoccupied states |k . Restrictions like energy, momentum, and spin conservations are taken into account by the particular choice of W (k, k ). With all terms included, the continuity equation in the phase space reads as

f

time

+

f

diff

+

f

 

field

=

f

scatter .

t

 

t

t

t

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

We can now write the

linearized

Boltzmann

equation as

 

 

 

 

f

 

 

e

 

1

 

f

scatter .

 

0 =

 

 

+ vk · r f

 

E +

 

vk × H · k f

 

(5.2.7)

 

t

h

c

t

 

 

 

 

¯

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

This form of Boltzmann’s equation is valid within the framework of linear response theory. Because we have retained only the gradient term r f , the theory is appropriate in the small q limit, but has no further limitations.

Next we solve the Boltzmann equation for a free gas of electrons in the absence of an external magnetic field (H = 0), with the electrons obeying quantum statistics and subjected only to scattering which leads to changes in their momenta. Let us assume that the distribution function varies only slightly from its equilibrium state fk0, and fk = fk0 + fk1, where

f 0

=

f 0(

Ek

)

exp

Ek EF

 

+

1

 

1

(5.2.8)

 

kBT

 

k

 

 

=

 

 

 

is the Fermi–Dirac distribution function. scattering term reduces to

fk

scatter

=

 

fk1 fk1

t

 

 

 

 

 

 

 

 

 

 

 

 

 

 

In this limit of small perturbation the

W (k, k ) dk .

If we make the assumption that the system relaxes exponentially to its equilibrium state after the perturbation is switched off (the relaxation time approximation), the distribution function has the time dependence fk1(t) = fk1(0) exp{−t}; therefore

5.2 Boltzmann’s transport theory

109

we find that the (k independent) scattering term is given by

 

fk

scatter

1

 

(5.2.9)

 

t

= − τ fk1 .

 

 

 

 

 

 

 

 

 

 

 

 

 

 

If the momentum change is small compared with the Fermi momentum (k

 

kF),

scattering across the Fermi surface can be neglected. Since

f

f

E

 

 

1k =

 

and (again

assuming H

 

0) the carrier velocity v

 

 

 

 

 

h

E k

 

 

 

=

k

= k

ω(k)

=

kE

(k) (which is the

 

 

 

 

 

 

¯

 

 

 

 

 

 

group velocity of the electrons), we can write Eq. (5.2.7) as

 

 

 

 

 

 

 

 

fk1

1

 

 

 

fk0

 

fk1

 

 

 

 

 

 

 

 

 

= vk · r fk eE · vk

 

+

 

.

 

 

(5.2.10)

 

 

t

∂E

τ

 

 

We can solve this equation by assuming that the distribution function follows the spatial and time dependent perturbation of exp{i(q · r ωt)} and has itself the form fk1(t) exp{i(q · r ωt)}. Taking the Fourier components in r and t, we solve Eq. (5.2.10) for f 1, obtaining

 

 

 

 

 

 

0

 

 

 

 

 

eE(q, ω) · vk(

f

.

 

f 1(q, k, ω)

=

E

(5.2.11)

 

 

1 i

ωτ

+ ivk · q

τ

 

 

 

 

 

 

The linearized distribution function yields a finite relaxation time τ and, consequently a finite mean free path = vFτ . The current density can in general be written as

J(q, ω) = −

2e

 

f 1(q, k, ω) vk dk

(5.2.12)

(2π )3

in terms of the distribution function f 1(q, k, ω), where vk is the velocity of the electrons for a given wavevector k and the factor of 2 takes both spin directions into account. Using Eq. (5.2.11) for f 1(q, k, ω) we obtain

 

2e2

 

 

τ E(q, ω) · vk(

f 0

)

 

 

 

 

 

 

J(q, ω) =

 

dk

1 iωτ + ivk · qEτ

vk .

(5.2.13)

(2π )3

As J(q, ω) = σˆ (q, ω)E(q, ω), the wavevector and frequency dependent conductivity can now be calculated. Let us simplify the expression first; for cubic symmetry

 

 

2e2

 

τ (nE · vk)vk(

f 0

)

 

 

ˆ

=

,

(5.2.14)

(2π )3

1 iωτ + ivk · qτ

 

σ (q, ω)

 

 

dk

 

E

where nE represents the unit vector along the direction of the electric field E. In the spirit of the Sommerfeld theory the derivative (f 0/∂E) is taken at E = EF, as only the states near the Fermi energy contribute to the conductivity. Second,

110

5 Metals

we consider this equation for metals in the limit T = 0, then the Fermi–Dirac distribution (5.2.8) is a step function at EF. With

T 0

f 0

=

 

{E EF}

(5.2.15)

∂E

 

lim

 

 

 

δ

 

 

the integral has contributions only at the Fermi surface SF. Under such circumstances we can reduce Eq. (5.2.14) to

 

2e2

 

 

τ (nE

vk)vk

 

 

f 0

 

dS

σˆ (q, ω) =

 

 

1

 

iωτ

·

ivk

 

qτ

 

 

 

 

 

dE

(2π )3

+

·

 

E

hvk

=

 

 

 

 

 

 

 

 

 

¯

 

(2π )3 E=EF

1 iωτ + ivk · qτ hvk

 

 

 

 

 

 

2e2

 

 

τ (nE · vk)vk

 

 

dSF

 

,

(5.2.16)

¯

where we have also converted the volume integral dk into one over surfaces dS of constant energy as

dk

=

dS dk

=

dS

dE

=

dS

dE

,

 

 

(5.2.17)

| kE|

hvk

 

 

 

 

 

 

 

 

¯

 

h2

2

 

and we have assumed a parabolic energy dispersion E(k) =

 

.

 

2¯m

k

 

5.2.2 The q = 0 limit

In the so-called local limit q 0, the conductivity simplifies to

σ (0, ω)

=

 

e2

 

τ (nE · vk)vk

 

1

dS

σ

 

1

;

(5.2.18)

4π 3h

vk

1 iωτ

dc 1 iωτ

ˆ

F =

 

 

 

 

¯

 

 

 

 

 

 

 

 

 

 

and we recover the familiar equation of the Drude model for the frequency dependent conductivity. The velocity term (nE ·vk)vk/vk averaged over the Fermi surface is simply 13 vF and therefore

 

=

e2

τ v2

=

 

e2τ 8π hk3

=

N e2

 

 

8π 3h

 

 

vF

 

 

4π 3h 3m

 

m

 

σdc

 

 

 

dSF

 

F

 

 

 

 

¯

F

 

 

τ

 

¯

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

EF dS = 2(4π kF2) and the

by considering the spherical Fermi surface for which

density of charge carriers N

3

/(

3

π 2)

. With our

prior assumption of quantum

= kF

 

 

 

 

!

 

 

statistics, only electrons near the Fermi energy EF are important, and therefore the mean free path = vFτ .

5.2.3 Small q limit

Next we explore the response to long wavelength excitations and zero temperature starting from Eq. (5.2.16). With the dc conductivity σdc = N e2τ /m, straightfor-

 

 

 

 

 

 

 

5.2

Boltzmann’s transport theory

 

 

111

ward integration yields2

 

 

 

 

 

σ (q, ω)

=

3σdc i

2

ω + i

1 + i/τ )2/(qvF)2

Ln

ω qvF + i

;

 

 

 

 

q2vF2

 

 

ˆ

 

4 τ

 

qvF

ω + qvF + i

 

 

 

 

 

 

 

 

 

 

 

(5.2.19)

where we recall the definition of the logarithm (Ln) of a complex value xˆ = |xˆ| exp{iφ} as the principal value

 

 

 

Ln{xˆ} = ln{|xˆ|} + iφ

 

(5.2.20)

with π < φ π . The first terms of the expansion

 

 

Ln{(zˆ + 1)/(zˆ − 1)} = 2(1/zˆ + 1/3zˆ2 + 1/5zˆ5 + · · ·)

 

give

 

 

 

 

 

 

 

 

σˆ (q, ω)

m

 

1 iωτ

1 5

1 iωτ

2

(5.2.21)

 

+ · · ·

 

N e2τ

1

1

 

qvFτ

 

 

for qvF < |ω + i|. This is called the homogeneous limit, the name referring to the case when the variation of the wavevector q is small. The second term in the square brackets is neglected for q = 0 and we recover the Drude form. An expansion in terms of + i/τ )/(qvF) leads to

ˆ

 

3π N e2

 

 

 

ω2

 

4ω

 

 

4qvFτ m

 

q2vF2 +

 

π qvF

σ (q, ω)

 

 

 

1

 

 

 

i

 

(5.2.22)

for |ω + i| < qvF,

in the so-called quasi-static limit.

The name refers to

the fact that in this limit the phase velocity of the electromagnetic field ω/q is small compared with the velocity of the particles vF. If the relaxation rate goes to zero (τ → ∞), or more generally for ωτ 1, the real and imaginary parts of the conductivity can be derived without having to limit ourselves to the long wavelength fluctuations. By utilizing the general transformation

lim

0 Ln{−|x| + i

} = ln |x| + i

π ,

(5.2.23)

1

 

 

 

 

 

 

 

 

2The integration (or summation) is first performed over the angle between k and q, leading to the logarithmic term

π

x

0

 

a + b cos x

and then from 0 to vF by using the relation

x ln{ax + b}dx = 2ba x

dx

 

1

ln

a

+ b

,

 

 

 

 

 

= b a

b

 

 

 

 

 

 

 

 

1 x2 + 1 x2 b2 ln{ax + b}.

4 2 a2

112 5 Metals

after some algebra we find that

 

 

 

 

 

 

 

 

3π

N e2

1

 

 

ω2

 

for

 

 

ω < qvF

 

 

lim σ

(

 

, ω)

 

 

 

 

4

mqvF

 

 

2

2

 

 

 

 

 

q

=

 

 

q

vF

 

 

 

 

 

 

 

 

 

 

 

(5.2.24a)

τ →∞ 1

 

 

0

 

 

 

 

 

for

 

 

ω > qvF

 

lim σ2

(q, ω)

 

 

3

 

N e2

 

 

2ω

 

 

 

 

 

ω2

 

1

ln

 

 

qvF ω

 

(5.2.24b)

=

 

4 mqvF

qvF

+ q2vF2

 

 

τ →∞

 

 

 

 

 

 

qvF

+

ω

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ω < qvF,

for both ω < qvF and ω > qvF. The conductivity σ1 is finite

only for

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

there are no losses in the opposite, homogeneous, limit. The reason for this is clear: ω > qvF would imply a wave with velocity ω/q greater than the Fermi velocity vF; the electron gas clearly cannot respond to a perturbation traveling with this velocity.

5.2.4 The Chambers formula

The q dependent response is intimately related to the non-local conduction where the current density J at the position r is determined also by fields at other locations r = r . Here we develop an approximate expression for the current which depends on the spatial distribution of the applied electric field. Such a situation may occur in the case of clean metals at low temperatures when the mean free path is large. Let us consider an electron moving from a point r to another position taken to be the origin of the coordinate system. At the initial location, the electron is subjected to an electric field E(r) which is different from that at the origin. However, because of collisions with the lattice or impurities, the momentum acquired by the electron from the field at r decays exponentially as the origin is approached. The characteristic decay length defines the mean free path , and the currents at the origin are the result of the fields E(r) within the radius of = vFτ .

The argument which accounts for such a non-local response is as follows. When an electron moves from a position (r dr) to r, it is influenced by an effective field E(r) exp{−r/ } for a time dr/vF. The momentum gained in the direction of motion is

dp(0) = −

e dr r

· E(r) exp{−r/ } ;

(5.2.25)

vF r

in the following we drop the indication of the position (0). By integrating the above equation from the origin to infinity, the total change in momentum for an electron at the origin is found. Performing this calculation for all directions allows us to map out the momentum surface in a non-uniform field, and the deviations from a sphere centered at the origin constitute a current J. The following arguments lead to the expression of the current. Only electrons residing in regions of momentum space not normally occupied when the applied field is zero contribute to the current.

5.2 Boltzmann’s transport theory

113

The density of electrons N moving in a solid angle d and occupying the net displaced volume in momentum space P is

N =

P

N =

(mvF)2 d dp

N =

3N d dp

,

(5.2.26)

P

4

π(mvF)3

4π vFm

 

 

 

3

 

 

 

 

 

where P is the total momentum space volume. The contribution to the current density from these electrons is

dJ = − N evF

r

 

3N e r

 

= −

 

 

 

d d p .

r

4π m

r

Substituting Eq. (5.2.25) into this equation and integrating over the currents given above yields

 

3σdc

 

r[r

 

E(r) exp

r/

]

 

 

J(r = 0) =

 

 

 

 

·

 

r4

{− }

 

dr

,

(5.2.27)

4π

 

 

 

 

 

since a volume element in

real

space

 

is

r2 dr d and

σdc

=

N e2τ /m =

N e2 /(mvF). Equation (5.2.27) represents the non-local generalization of Ohm’s law for free electrons, and reduces to J = σdcE for the special case where 0, as expected. The Chambers formula [Cha90, Pip54b] is valid for finite momentum, but as the Fermi momentum is not explicitly included its use is restricted to q < kF, and in general to the small q limit. A quantum mechanical derivation of Chambers’ result was given by Mattis and Dresselhaus [Mat58]. In Appendix E the non-local response is discussed in more detail.

5.2.5 Anomalous skin effect

In Section 2.2 we introduced the skin effect of a metal and found that the skin depth is given by

c

δ0(ω) = (2π ωσdc)1/2

in the low frequency, so-called Hagen–Rubens regime. In this regime we also obtained the following equation for the surface impedance

Zˆ S = RS + iXS =

(2π )2 δ0

(1 i) ,

(5.2.28)

c λ0

where the frequency dependence enters through the skin depth δ0 and through the free space wavelength λ0 of the electromagnetic field. The condition for the above results is that the mean free path is smaller than δ0 and hence local electrodynamics apply. At low temperatures in clean metals this condition is not met because becomes large, and the consequences of the non-local electrodynamics have to be

114

5 Metals

 

 

δeff

Fig. 5.9. Electron trajectories near the surface of a material (solid line) in the case of the anomalous regime, where the mean free path is much larger than the effective skin depth δeff (dashed line). According to the ineffectiveness concept, only electrons traveling nearly parallel to the surface are effective in absorbing and screening electromagnetic radiation when δeff, since the carriers moving perpendicular to the surface leave the skin layer before they scatter.

explored in detail. This can be done by using the Chambers formula (5.2.27), or alternatively the solution of the non-local electrodynamics, as in Appendix E.1.

An elegant argument, the so-called ineffectiveness concept due to Pippard [Pip47, Pip54a, Pip62] reproduces all the essential results. If the mean free path is larger than the skin depth, the effect on the electrons which move perpendicular to the surface is very different from those traveling parallel to the surface. Due to the long mean free path, electrons in the first case leave the skin depth layer without being scattered; the situation is illustrated in Fig. 5.9. Thus, only those electrons which are moving approximately parallel to the surface (i.e. their direction of motion falls within an angle ±γ δeff/ parallel to the surface of the metal) contribute to the absorption [Pip54b, Reu48]. The (reduced) number of effective electrons Neff = γ N δeff/ also changes the conductivity, which we write as:

δeff

 

(5.2.29)

σeff = γ σdc

,

where γ is a numerical factor of the order of unity, and can be evaluated by using Chamber’s expression [Abr72, Pip54b, Reu48]. The factor depends on the nature of the scattering of the electrons at the surface; for specular reflection γ = 8/9 and for diffuse reflection γ = 1. In turn the effective conductivity leads to a modified skin depth δeff = c/ (2π ωσeff)1/2, and by substituting this into Eq. (5.2.29) we obtain through selfconsistence for the effective skin depth

δeff =

c2

 

 

1/3

=

c2 mvF

 

1/3

 

 

 

 

 

 

2π ω γ σdc

 

2π ω γ N e2

 

,

(5.2.30)

and for the effective conductivity

 

 

 

 

 

 

 

 

σeff =

γ c σdc

 

2/3

=

γ c N e2

 

2/3

(5.2.31)

2π ω

 

 

2π ω mvF

.

 

 

 

 

 

 

 

5.3 Transverse response for arbitrary q values

 

115

We can

then

use

 

this

σ

eff for evaluating the optical

constants n + ik =

 

1/2

 

 

 

 

(4π iσeffω)

 

/c and the surface impedance Zˆ , utilizing Eq. (2.3.28). The end result

for the surface impedance is

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1 i

 

 

 

2π ω

 

2

 

 

1/3(1

 

i

 

)

 

 

 

 

 

 

 

 

 

 

 

 

3

 

 

 

 

 

 

Z

 

R

 

 

 

iX

 

 

 

 

 

 

 

 

 

 

 

3

 

=

 

+

S =

 

 

 

 

=

c2

 

 

γ σdc

 

ˆ S

 

S

 

 

δeffσeff

 

 

 

 

 

 

 

 

=

 

2π ω

 

2

 

mvF

1/3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(1 i3) .

 

 

 

 

(5.2.32)

 

 

 

 

c2

 

 

γ N e2

 

 

 

 

 

 

 

Both RS and XS increase with the ω2/3 power of frequency, and the ratio XS/RS = −3; both features are dramatically different from those obtained in the case of the normal skin effect. There is also another significant difference. The effective conductivity and surface resistance are independent of the mean free path or any other temperature dependent parameter for the anomalous skin effect. This can be used to explore the characteristics of the Fermi surface, as will be discussed in Chapter 12.

5.3 Transverse response for arbitrary q values

By employing Boltzmann’s equation we have limited ourselves to long wavelength fluctuations with the Drude–Sommerfeld model representing the q = 0 limit. This limit is set by performing an expansion in terms of q; however, more importantly, we have neglected the limits in the transition probability Wkk which are set by the existence of the Fermi surface separating occupied and unoccupied states. These have been included in the selfconsistent field approximation, which was discussed in Section 4.3. We now utilize this formalism when we evaluate σˆ (q, ω) for arbitrary q and ω parameters.

Let us first look at the condition for electron–hole excitations to occur; these will determine the absorption of electromagnetic radiation. In the ground state, such pair excitations with momentum q are allowed only if the state k is occupied and the state k = k + q is empty, as indicated in Fig. 5.10. The energy for creating such electron–hole pairs is

 

h2

h2

E = E(k + q) E(k) =

2¯m

(k + q)2 k2 =

2¯m

(q2 + 2k · q) . (5.3.1)

For any q the energy difference E and

Emax(q)

has a maximum value when q k and |k| = kF,

 

h2

 

 

 

 

=

¯

(q2

+

2qkF) .

(5.3.2)

2m

 

 

 

116

 

 

5 Metals

 

 

 

kF

 

q

q

q

(a)

(b)

 

(c)

Fig. 5.10. Fermi sphere with radius kF corresponding to electronic states with momentum h¯ k and h¯ (k + q). The shaded region represents the momentum space to which single- particle–hole pairs can be excited with momentum h¯ q so that |k| < |kF| and |k+ q| > |kF| (only values of k where k is occupied and k + q is empty, and vice versa, contribute): (a) small q values, (b) intermediate q values, (c) large momentum |q| > |2kF|.

The minimum excitation energy is zero for |q| ≤ 2kF. However, for |q| > 2kF the minimum energy occurs for q k and |k| = kF, and

 

 

h2

(q2 2qkF)

for |q| > 2kF

 

min(q)

2¯m

(5.3.3)

E

=

0

 

otherwise ,

 

 

 

 

 

 

 

finally leading to the following condition for electron–hole excitations to occur:

 

ω

q

 

< 1

 

ω

+

q

.

qvF

2kF

< qvF

2kF

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

This region is indicated in Fig. 5.11 by the hatched area.

Next we want to evaluate the conductivity for intraband transitions (l = l ) starting from Eq. (4.3.32). In this case,

σ (

, ω)

 

N e2

 

lim

i

 

e2

 

 

 

 

 

f 0( k)

 

 

 

= i

ωm

 

 

 

k

k

 

kq

Ehω

 

ihη

 

ˆ q

 

η0 ωm2

 

E

 

 

 

 

 

 

 

 

f 0(Ek)

 

 

E

 

− ¯

¯

 

 

 

 

k+q

 

ihη

| k + q|p|k |2

 

,

(5.3.4)

 

 

E

k hω

 

 

 

 

 

 

 

E − ¯

¯

 

 

 

 

 

 

 

 

if we consider one single band. Here we have to split up the summation as

k

k q Ek+k

hω

 

Ekihη =

k

 

 

k

 

k q

 

Ekhω ihη

 

f

q

f

(

)

 

 

 

 

 

 

f

(

 

 

)

 

 

 

 

 

E + E − ¯

− ¯

E E − ¯

− ¯

 

 

 

 

 

 

 

 

 

E +

 

 

 

f

(Ek)

 

 

 

.

 

 

 

 

 

 

 

 

E

 

 

¯

 

 

 

 

 

 

 

q

k

¯

ω

 

 

 

 

 

 

 

 

 

k

 

k

 

 

 

h

 

ihη

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

We have replaced k + q by k in the first term, which is allowed as the summation over k involves all the states, just as the summation over k+q does. We now replace the summation 1 #k by the integration over the k space 2/(2π )3 ! dk and, in the spirit of the relaxation time approximation, η is substituted by the relaxation

5.3 Transverse response for arbitrary q values

117

 

 

 

 

 

 

 

)

 

 

 

 

 

 

 

 

 

 

 

 

2qk

F

 

 

 

 

 

 

 

 

 

 

 

+

 

 

 

 

 

 

 

 

 

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

h

(q

 

 

Electron–hole

 

 

 

hω

 

 

2m

 

 

 

 

 

 

 

 

 

excitations

 

 

 

 

σ1 = 0

 

=

 

 

 

 

 

 

transfer

ω

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

h

 

 

 

 

 

 

 

 

 

 

)

 

 

 

 

 

 

 

 

 

 

 

2qk

F

 

 

 

 

 

 

 

 

 

 

 

 

Energy

 

 

 

 

 

 

 

 

 

 

 

qvF < ω

 

 

 

σ1 0

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

h

(q

 

 

 

 

 

 

 

 

 

 

 

2m

 

 

 

 

ωp(q)

 

 

 

 

 

 

 

 

 

 

hωp

h2

 

 

 

 

 

hω

=

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

hω =

2m (2qkF q2)

 

 

 

 

 

00

qvF > ω

 

 

 

 

 

 

σ1 = 0

 

 

 

 

 

 

2kF

 

 

 

 

 

 

Wavevector q

Fig. 5.11. Excitation spectrum of a three-dimensional free-electron gas; the transferred energy is plotted as a function of transferred momentum. The pair excitations fall within the shaded area. In the region h¯ ω > (h¯ 2/2m)(q + 2kF)q, the absorption vanishes since h¯ ω is larger than energies possible for pair creation. For h¯ ω < (h¯ 2/2m)(q 2kF)q we find σ1 = 0. Also shown is the dispersion of the plasma frequency ωp(q) calculated from

Eq. (5.4.28) using typical values for sodium: ωp = 5.9 eV and vF = 1.1 × 108 cm s1.

rate 1. Evaluating the matrix element for the conduction band in the case of a free-electron gas Ek = h¯22mk2 , we can rewrite the diagonal elements of Eq. (5.3.4) as [Lin54]

σ (q, ω)

 

i

e2

 

2 f 0(

 

)

 

k2

 

k · q

2

 

 

 

 

 

1

 

 

 

 

=

mω

 

Ek

 

q

q2

 

2k q

 

 

i/τ )

ˆ

 

k

 

 

 

 

 

 

2m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+

·

h

+

 

 

 

 

 

 

 

 

 

 

 

+ i/τ ) +

2

 

 

¯

(5.3.5)

 

 

+ q2 2k · q + 2hm

.

 

 

 

 

 

 

 

 

 

 

 

 

1

 

 

 

 

 

 

1

 

 

 

 

 

 

 

 

 

¯

The first square bracket is the matrix element | k + q|p|k |2 evaluated using Eq. (3.1.5) to consider the transverse components. For transverse coupling between the vector potential A and momentum p, we keep only those components of p